Solvent Reorganization as a Governing Factor in the Kinetics of

aRX•- is the radius of the sphere approximating the region of the starting anion radical where the charge is initially located and aR•,X-, the rad...
0 downloads 0 Views 99KB Size
9788

J. Am. Chem. Soc. 1996, 118, 9788-9789

Solvent Reorganization as a Governing Factor in the Kinetics of Intramolecular Dissociative Electron Transfers. Cleavage of Anion Radicals of r-Substituted Acetophenones Claude P. Andrieux,1a Jean-Michel Save´ant,*,1a Andre´ Tallec,1b Robert Tardivel,1b and Caroline Tardy1a Laboratoire d’Electrochimie Mole´ culaire de l’UniVersite´ Denis Diderot, Unite´ Associe´ e au CNRS No. 438, 2 place Jussieu, 75251 Paris Cedex 05 Laboratoire d’Electrochimie de l’UniVersite´ de Rennes Unite´ Associe´ e au CNRS No. 439, Campus de Beaulieu 35042 Rennes Cedex, France ReceiVed July 3, 1996 Cleavage of ion radicals2-6 may be viewed as an intramolecular electron transfer coupled with bond breaking in cases where the unpaired electron initially resides in an orbital that does not belong to the leaving group:7 -•

R-X h R + X •

-

The unpaired electron jumps from the initial orbital to the σ* orbital of the R-X bond which breaks simultaneously. Hence, a model of the dynamics of such reactions has been proposed8 where the two main reaction coordinates are the bond distance and a fictitious charge representing solvent reorganization in the Marcus way.9 The activation free energy, ∆Gq, is thus related quadratically to the standard free energy of the reaction, ∆G°:

∆Gq ) ∆G0q(1 + ∆G°/4∆G0q)2

(1)

The standard free energy of activation, ∆G0q, is the sum of two contributions, one related to bond breaking and the other to solvent reorganization: (1) (a) Universite´ Denis Diderot. (b) Universite´ de Rennes. (2) (a) Neta, P.; Behar, D. J. Am. Chem. Soc. 1980, 102, 4798. (b) Behar, D.; Neta, P. J. Phys. Chem. 1981, 85, 690. (c) Behar, D.; Neta, P. J. Am. Chem. Soc. 1981, 103, 103. (d) Behar, D.; Neta, P. J. Am. Chem. Soc. 1981, 103, 2280. (e) Bays, J. P.; Blumer, S. T.; Baral-Tosh, S.; Behar, D.; Neta, P. J. Am. Chem. Soc. 1983, 105, 320. (f) Norris, R. K.; Barker, S. D.; Neta, P. J. Am. Chem. Soc. 1984, 106, 3140. (g) Meot-Ner, M.; Neta, P.; Norris, R. K.; Wilson, K. J. Phys. Chem. 1986, 90, 168. (3) (a) Saeva, F. D. Top. Curr. Chem. 1990, 156, 61. (b) Saeva, F. D. Intramolecular Photochemical Electron Transfer (PET) - Induced Bond Cleavage Reactions in some Sulfonium Salts Derivatives. In AdVances in Electron Transfer Chemistry; Mariano, P. S., Ed.; JAI Press: New York, 1994; Vol. 4, pp 1-25. (c) Arnold, B. R.; Scaiano, J. C.; McGimpsey, W. G. J. Am. Chem. Soc. 1992, 114, 9978. (d) Chen, L.; Farahat, M. S.; Gan, H.; Farid, S.; Whitten, D. G. J. Am. Chem. Soc. 1995, 117, 6399. (4) (a) Tanner, D. D.; Chen, J. J.; Chen, L.; Luelo, C. J. Am. Chem. Soc. 1991, 113, 8074. (b) Tanko; J. M.; Drumright, R. E.; Sulemen, N.; Brammer, L. E. J. Am. Chem. Soc. 1994, 116, 1585. (c) Andersen, M. L.; Mathivanan, N.; Wayner, D. D. M. J. Am. Chem. Soc. 1996, 118, 4871. (5) Severin, M. G.; Arevalo, M. C.; Maran, F.; Vianello, E. J. Phys. Chem. 1993, 97, 150. (6) (a) Save´ant, J.-M. Single Electron Transfer and Nucleophilic Substitution. In AdVances in Physical Organic Chemistry; Bethel, D., Ed.; Academic Press: New York, 1990; Vol. 26, pp 1-130. (b) Save´ant, J.-M. Acc. Chem. Res. 1993, 26, 455. (c) Save´ant, J.-M. Dissociative Electron Transfer. In AdVances in Electron Transfer Chemistry; Mariano, P. S., Ed.; JAI Press: New York, 1994; Vol. 4, pp 53-116. (d) Save´ant, J.-M. Tetrahedron 1994, 50, 10117. (7) In the opposite case,7b-j the reaction is better viewed as a homolysis and the intramolecular (small) reorganization pertains essentially to rehybridization.7b The solvent reorganization energy is also expected to be small since the charge stays on the same portion of the molecule during bond cleavage. (b) Anne, A.; Fraoua, S.; Moiroux, J.; Save´ant, J.-M. J. Am. Chem. Soc. 1996, 118, 3938. (c) Maslak, P.; Vallombroso, T. M.; Chapman, W. H.; Narvaez, J. N. Angew. Chem., Int. Ed. Engl. 1994, 33, 73. (d) Maslak, P.; Narvaez, J. N.; Vallombroso, T. M. J. Am. Chem. Soc. 1995, 117, 12373. (e) Maslak, P.; Chapman, W. H.; Vallombroso, T. M.; Watson, B. A. J. Am. Chem. Soc. 1995, 117, 12380. (8) Save´ant, J.-M. J. Phys. Chem. 1994, 98, 3716. (9) Marcus, R. A. J. Chem. Phys. 1965, 43, 679.

S0002-7863(96)02256-1 CCC: $12.00

∆G0q )

DRX•- + λ° 4

(2)

where the first term may be derived from accessible molecular parameters according to

DRX•- ) DRX + E°RX/RX•- - E°R•/R- + T(SR• - SRX) + T(SRX•- - SR-) (3) (the D, E°, and S values are the bond dissociation energies, the standard potentials, and the molar entropies of the subscript species, respectively). The solvent reorganization energy arises from the fact that, during the reaction, the charge moves from one portion of the molecule to another. λ° may thus be estimated from

(

λ° ) e02

)(

)

1 1 1 1 1 + 2aRX•- 2aR•,X- d Dop Ds

(4)

aRX•- is the radius of the sphere approximating the region of the starting anion radical where the charge is initially located and aR•,X-, the radius of the sphere approximating the location of the charge on the leaving group when the bond is broken. d is the distance between the centers of charge at the transition state. The D values are the optical and static dielectric constants, respectively. The standard free energy of anion radical cleavage may be obtained from eq 5, where ∆SRX is the entropy for the cleavage of R-X.

∆G° ) DRX + E°RX/RX•- - E°X•/X- - T∆SRX

(5)

So far, only the influence of the intramolecular factors (through eqs 1-3) have been investigated experimentally (see refs 6d and 8) in reactions where they are largely predominant over possible solvent reorganization effects.10 In order to maximize the effect of solvent reorganization and minimize the contribution of intramolecular factors, we have selected a series of anion radicals generated by electrochemical reduction of the following R-substituted acetophenones. Z

CH2X O

X Br

Z 4-NO2

1a

4-CN

1b

O C Ph

H

2a

O

4-NO2

2b

4-OCH3

2c

X OCH3

Z

H

O

3a 3b

4-OCH3 SC2H5

H

4a

Over the whole series of anion radicals generated by cyclic voltammetry, only 2b•- was found to be stable down to the lowest scan rate (0.1 V/s).11 4a exhibts an irreversible wave at low scan rate which becomes reversible around 10 V/s. The cleavage rate constant and the standard potential for the formation of the anion radical of 4a were derived from these cyclic voltammetric data (see the supporting information). All the other compounds give rise to an irreversible wave between (10) (a) The influence of orbital symmetry restrictions has also been examined.10b (b) Adcock, W.; Andrieux, C. P.; Clark, C. I.; Neudeck, A.; Save´ant, J.-M.; Tardy, C. J. Am. Chem. Soc. 1995, 117, 8285. (11) It is interesting to note that charge transfer to 2b is relatively slow. The standard rate constant, kS ) 0.22 cm-1 s-1, was determined from cyclic voltammograms (see the supporting information). Charge transfer to acetophenone itself is even slower (kS ) 0.14 cm-1 s-1), pointing to a strong localization of the negative charge on the carbonyl oxygen of the anion radical in both cases, albeit somewhat less with 2b than with acetophenone.

© 1996 American Chemical Society

Communications to the Editor

J. Am. Chem. Soc., Vol. 118, No. 40, 1996 9789

Table 1. Reactivity Dataa contributions cmpd 1a 1b 2a 2c 3a 3b 4a

log k (s-1) 6.6 7.8 6.4 7.6 5.4 5.4 1.2

DRXb c

1.85 1.90c 2.53d 2.53d 1.78d 1.78d 2.14d

E°X•/Xe

1.44 1.44 e 1.24f,g 1.24f,g 0.24f,h 0.24f,h 0.06f,i

-E°RX/RX•-

-∆G° j (cleav)

∆G0q k

intral

solvm

0.83 1.04 1.81 1.97 1.81 1.88 1.87

0.71 0.87 0.82 0.98 0.56 0.63 0.08

0.65 0.66 0.72 0.7 0.68 0.69 0.71

0.26 0.22 0.18 0.14 0.00 0.00 0.07

0.39 0.44 0.54 0.56 0.68 0.69 0.64

a Energies in eV, potentials in V vs SCE. b From the value DRX ) 2.05 eV obtained from the peak potential, Ep, of the dissociative electron transfer reduction of PhCOCH2Br by means of the equation: DRX ) (2/3)(E°X•/X- - Ep)14a,b (details on the electrochemistry of PhCOCH2Br will be published elsewhere). c The value for PhCOCH2Br is corrected for the weakening of the bond by the para electron-withdrawing substituent as for the corresponding benzyl bromides.14a,c d From DPhCOCH2X ) DPhCOCH2Br + DPhCH2X - DPhCH2Br; the DPhCH2X values are obtained from ref 15. e From ref 14a,b. f Assumed to be the same as in acetonitrile. g See ref 16. h From ref 17. i From ref 16b. j From the application of eq 5, taking ∆S ) 1 meV K-1 which corresponds to the formation of two molecules from one and is small and does not vary significantly in the series. k The preexponential factor is taken equal to kT/h ) 5 × 1012 M-1 s-1. l In the application of eq 3, E°R•/R- was taken as equal to 0.00 V vs SCE18 and the molar entropies of the charged species (which are the predominant entropic terms) were assumed to compensate each other. m From the difference between the exprimental intrinsic barrier and the contribution of bond breaking (intra).

0.1 and 100 V/s. The values of the cleavage rate constants and of the standard potential for the formation of the anion radical were derived from the variation of the peak potentials with scan rate and from the value of the peak width12 according to a procedure described in the supporting information. The results (Table 1)13 are displayed in Figure 1 as a plot of log k (cleavage) vs the standard free energy of the reaction. The latter quantity was estimated as described in Table 1. The standard potentials E°RX/RX•- were obtained from the same experiments (see the supporting information). There is a clear correlation between the rate constants and the standard free energy of the reaction of the type depicted by eq 1 with a value of the average symmetry factor close to 0.5 (0.45). The average intrinsic barrier is 0.70 eV. Why is the intrinsic barrier so large, and why does it depend so weakly upon molecular structure? We may estimate the contribution of the intramolecular bond cleavage by application of eq 3. As seen in Table 1, it represents (12) (a) Nadjo, L.; Save´ant, J.-M. J. Electroanal. Chem. 1973, 48, 113. (b) Andrieux, C. P.; Save´ant, J.-M. In Electrochemical Reactions in InVestigation of Rates and Mechanisms of Reactions, Techniques of Chemistry; Bernasconi, C. F., Ed.; Wiley: New York, 1986; Vol. VI/4E, Part 2, pp 305-390. (13) The value found for 3a is in good agreement with the values recently reported4c for an extended series of R-phenoxyacetophenones bearing various substituents on the phenoxy ring, although not a 3-methyl group as in 3a. (14) (a) Andrieux, C. P.; Le Gorande, A.; Save´ant, J.-M. J. Am. Chem. Soc. 1992, 114, 6892. (b) Andrieux, C. P.; Differding, E.; Robert, M.; Save´ant, J.-M. J. Am. Chem. Soc. 1993, 115, 6592. (c) Clark, K. B.; Wayner, D. D. M. J. Am. Chem. Soc. 1991, 113, 9363. (15) (a) Benson, S. N. Thermochemical kinetics; Wiley: New York, 1976. (b) Handbook of Chemistry and Physics, 72nd ed.; CRC: Cleveland, 19911992; pp 9, 116-121. (16) (a) A reliable value, 1.06 V vs SCE, is available for CH3CO2-,16b but not for PhCO2-. The value given in ref 16, 0.36 V vs SCE, is smaller than for CH3CO2-, which seems counterintuitive in view of the fact that CH3CO2- is a stronger base than PhCO2- (the pKa values in acetonitrile are 22.3 and 20.7, respectively16c,d). Starting from the E° value of CH3CO2-, we estimated the value for PhCO2- from the equation

E°PhCO2- - E°CH3CO2- (V) ) DPhCO2-H - DCH3CO2-H + 0.06(pKa,PhCO2H - pKa,CH3CO2H) assuming that the entropic terms, ∆SRX/R•+H•, are about the same in both cases. (b) Eberson, L. Acta Chem. Scand., B 1984, 439. (c) Kolthoff, I. M.; Chantoni, M. K.; Bhownik, S. J. Am. Chem. Soc. 1968, 90, 23. (d) Coetzee, J. F. Prog. Phys. Org. Chem. 1967, 4, 76. (17) Hapiot, P.; Pinson, J.; Yousfi, N. New J. Chem. 1992, 16, 877. (18) (a) The peak potential of PhCOCH2- in DMSO is -0.10 V vs SCE at 0.1 V/s.18b The negative shift caused by a diffusion-controlled followup dimerization may be estimated as 90-100 mV, leading to an E° value of 0.00.18b,c (b). Bordwell, F. G.; Harrelson, J. A. Can. J. Chem. 1990, 68, 1714. (c) Andrieux, C. P.; Hapiot, P.; Pinson, J.; Save´ant, J.-M. J. Am. Chem. Soc. 1993, 115, 7783.

Figure 1. Variation of the cleavage rate constant with the standard free energy of the reaction.

a modest-to-negligible percentage of the total barrier, particularly for compounds that do not bear ring substituents. We are thus led to conclude that solvent reorganization is the main factor dictating the height of the intrinsic barrier, contributing 75-100% to the total in the acetophenone series. This observation falls in line with the fact that, in the initial state of the anion radical, the solvent is organized around a negative charge which is mostly concentrated on the carbonyl oxygen and has to reorganize around the negative charge which develops on the leaving group during the reaction. Such values of the solvent reorganization energy are compatible with reasonable values of the two solvation spheres’ radii, on the order of 2 Å each, and of the distance between the two centers of charge, on the order of 6 Å. A radius of 2 Å for the solvation sphere of the oxygen carbonyl is also compatible with the small value, 0.14 cm/s, found for the standard rate constant of the electrochemical reduction of acetophenone. Obviously, these estimations are very approximate. Nevertheless, it is noteworthy that the contribution of solvent reorganization to the intrinsic barrier appears to be less with 1a and 1b, where charge concentration on the oxygen carbonyl is expected to be less than with the other compounds owing to charge delocalization on the 4-NO2 and 4-CN groups, respectively. Supporting Information Available: Cyclic voltammetric data and determination of the cleavage rate constant and standard potential for the formation of the anion radicals from these data (4 pages). See any current masthead page for ordering and Internet access instructions. JA9622564